Electrical Characterization of Bismuth Sulfide ... - ACS Publications

Abstract. A new method for determining the resistivity of templated Bi2S3 nanowires by conductive atomic force (C-AFM) microscopy is described in this...
0 downloads 0 Views 1MB Size
19680

J. Phys. Chem. C 2008, 112, 19680–19685

Electrical Characterization of Bismuth Sulfide Nanowire Arrays by Conductive Atomic Force Microscopy Pavels Birjukovs,† Nikolay Petkov,‡,§ Ju Xu,‡ Janis Svirksts,| John J. Boland,§ Justin D. Holmes,‡,§ and Donats Erts*,† Institute of Chemical Physics and Department of Chemistry, UniVersity of LatVia, Raina BlVd. 19, LV1586 Riga, LatVia, Materials and Supercritical Fluids Group, Department of Chemistry and the Tyndall National Institute, UniVersity College Cork, Cork, Ireland, and Centre for Research on AdaptiVe Nanostructures and NanodeVices (CRANN), Trinity College Dublin, Dublin 2, Ireland ReceiVed: June 19, 2008; ReVised Manuscript ReceiVed: October 14, 2008

A new method for determining the resistivity of templated Bi2S3 nanowires by conductive atomic force (CAFM) microscopy is described in this paper. Unlike other vertical C-AFM approaches, in our method, resistance measurements were carried out along the lengths of the nanowires. Nanowires embedded within anodic alumina membranes were exposed for contact by etching away the alumina template to form an open array of parallel nanowires. From these measurements, the contact resistance between the gold electrodes and the C-AFM probe could be determined and subtracted to give the intrinsic resistivity of the nanowires. The resistivity of the nanowires determined in such a horizontal configuration was 10-100 times lower than the resistivity determined when the same nanowires were contacted in a vertical configuration. Introduction Semiconductor nanowires are promising candidates as building blocks for the bottom-up assembly of nanosized devices and circuits.1 Individual semiconductor nanowires have been exploited for a number of nanoscale applications that include high-speed field effect transistors (FETs),2-4 light emitting devices,5 single electron memory devices,6 nanoelectromechanical devices,7 chemical and biological sensors,8 and logic circuits.9 However, the successful incorporation of nanowires into optical and electronic devices will require their assembly into electrically addressable high-density architectures so that their unique transport properties, either individually or collectively, can be utilized. Several research groups have reported the synthesis of ordered arrays of semiconducting nanowires, such as within the pores of ordered templates,10-14 where nanowire densities of up to 2 × 1012 cm-2 can be achieved.15 Much work has focused on the deposition of ordered arrays of nanowires, for example, ZnO,16 Ge,17-20 Si,21-23 and Bi2S313 inside anodized aluminum oxide (AAO) membranes. However, finding reliable methods for the fast screening and electrical characterization of individual nanowires within these arrays is a challenge. In previous reports, nanowires have been characterized as groups of nanowires by using macrocontacts on both sides of a templating membrane,16,18,19,21-23 or individually by liberating nanowires from the template and positioning a single nanowire between metal contacts. We have developed methods for determining the conductive properties of individual nanowires within a matrix using conductive atomic force microscopy (C-AFM).17,18,20 In this approach, the nanowires are contacted in their vertical orientation within the membrane by depositing a macrocontact on one side of the membrane and contacting * To whom correspondence should be addressed. Telephone: +371 67033875. Fax: +371 7033874. E-mail: [email protected]. † Institute of Chemical Physics, University of Latvia. ‡ University College Cork. § Trinity College Dublin. | Department of Chemistry, University of Latvia.

the exposed ends of the nanowire at the other end of the membrane using a C-AFM tip. With this methodology, both topography and conductivity maps can be obtained simultaneously, giving evidence of an ordered arrangement of conductive nanostructures. We have also used the vertical C-AFM method to determine whether nanowires run continuously throughout the membrane, from top to bottom. Nonetheless, all of the C-AFM measurements undertaken in the vertical configuration are two-point contact measurements, and a lot of effort is required to eliminate the contact resistance contribution, such as using different sized contact pads or thinning of the nanowire-template samples. In this paper, we describe the use of polishing and etching techniques to access arrays of Bi2S3 nanowires, in their longitudinal direction, and employ C-AFM techniques to determine the resistance along the full lengths of the nanowires. We are interested in studying Bi2S3 nanowires, as they have numerous potential applications as a material for photovoltaic24,25 and thermoelectric26 devices. However, the horizontal C-AFM technique described in this paper could be applied to other nanowire materials. Using this new technique, we have been able to determine the intrinsic resistivity of the nanowires of single crystalline, highly oriented nanowires within a templating matrix. The nanowire resistivity data obtained using the horizontal C-AFM technique were compared to those obtained using the vertical C-AFM method. Experimental Section Porous AAO membranes were purchased from Whatman, under the commercial name Anodisc filters, with a nominal pore size of about 200 nm. Porous AAO membranes with a pore diameter of 80 nm were synthesized in-house using a wellknown procedure described previously.27,28 A single source precursor, bismuth bis(diethyldithiocarbamate) [Bi(S2CNEt2)3] complex, was prepared by the reaction of Bi2O3 with carbon disulfide and diethylamine in methanol and was thoroughly purified by repeated recrystallization.13 The

10.1021/jp805422k CCC: $40.75  2008 American Chemical Society Published on Web 11/14/2008

Characterization of Bi2S3 Nanowire Arrays

Figure 1. SEM images of (a) perpendicular and (b) tilted sample surfaces after being etched with 9% H3PO4 for 40 min.

synthesis of arrays of single crystal Bi2S3 nanowires was carried out in a flat bottom glass tube (inner diameter 14 cm). In a typical experiment, the organometallic precursor (∼300 mg) was placed on the surface of the AAO membrane in the glass tube and heated to 200 °C under vacuum. After the precursor melted, the temperature was maintained for another 30 min. The vacuum was then disconnected, and the glass tube containing the sample was placed onto a hot plate at 300 °C for 30 min until the Bi2S3 precursor had completely decomposed. The crystal structure of the Bi2S3 nanowires was examined by selected area electron diffraction (SAED) and high-resolution transmission electron microscopy (HRTEM) combined with fast Fourier transforms (FFTs) on a JEOL 2010 HRTEM apparatus equipped with an Oxford Instrument EDX system. X-ray diffraction (XRD) data were collected on a Phillips Xpert PRO diffractometer using Cu KR radiation, an anode current of 40 mA, and an accelerating voltage of 40 kV. Data were collected in a range of 10-90 2Θ degrees with a step size of 0.005° and a scan rate of 0.2 s per step. The AAO samples were polished, etched, and then placed on a glass slide that minimized incoherent background scattering contributions. In order to access the electrical properties of the nanowires, the AAO membranes containing the Bi2S3 nanowires were treated in a way similar to that described previously18 prior to the AFM measurements. The surfaces of the AAO membranes were mechanically polished with diamond paste with grain sizes of 1 µm, 250 nm, and 50 nm in order to remove any Bi2S3 film from both sides of the membrane. Additionally the bottom surface was chemically etched with 9% H3PO4. Scanning electron microscopy (SEM) images of the membrane surface before and after etching showed that the ends of the nanowires were liberated from the membrane, exposing their tips (Figure 1). Before the gold macroelectrode was deposited onto the bottom of a membrane, the surface was cleaned with Ar ions at an energy of 5 keV and a current of 120 µA for 5-10 s, using a precision etching coating system (Gatan), to remove contamination. The surface of the membrane was then coated with a 50 nm thick layer of Au metal film in the same chamber without breaking the vacuum at 10-6 Torr and a coating speed of 1-1.5 Å s-1. To perform resistance measurements in a horizontal configuration (along the longitudinal axis of the nanowires), the membranes, with the deposited bottom electrode, were glued inside an epoxy in between two glass slides. The samples were then mechanically polished to expose the nanowires in the sample plane. Afterward, polishing the surface of the samples was chemically etched with 9% H3PO4. The surfaces of the membranes and the contacts were visualized using a Hitachi S4800 scanning electron microscope. An atomic force microscope (MFP3DTM, Asylum Research) in contact mode was used for topography and current mapping. Conductive Pt-coated silicon atomic force microscopy (AFM)

J. Phys. Chem. C, Vol. 112, No. 49, 2008 19681

Figure 2. Schematics of conductivity measurements through individual nanowires by C-AFM: (a) vertical configuration where the AFM probe is contacting the ends of nanowires and (b) horizontal orientation where the AFM probe is contacting the nanowire along the longitudinal nanowire axis.

Figure 3. (top left) TEM image of Bi2S3 nanowires isolated from 80 nm membranes, (top right) TEM image of 80 nm single Bi2S3 nanowire, and (bottom) HRTEM image of the same nanowire taken within the marked area with the corresponding growth direction shown with white arrows (inset: corresponding FFTs).

tips (NTMDT CSC12/Pt) and boron-doped diamond tips (NTMDT DCP20) were used in the experiments described in this paper. Results and Discussion The electrical properties of the individual Bi2S3 nanowires within the AAO matrix were analyzed in two different configurations throughout the study: (i) with the conductive AFM probe in contact with the ends of individual nanowires oriented vertically (Figure 2a) and (ii) with a C-AFM probe in contact with individual nanowires along their longitudinal axis (Figure 2b). The growth, structure, and encapsulation of the Bi2S3 nanowires within the pores of the AAO membranes were determined by XRD, HRTEM, and SAED. A detailed description of the crystallography and growth mechanism of the Bi2S3 nanowire arrays within the AAO membranes is presented elsewhere.13 Briefly, highly oriented single crystalline arrays of Bi2S3 nanowires within AAO membranes were obtained by growing them exclusively along the [002] growth direction. Figure 3 shows TEM images of Bi2S3 nanowires that have been isolated from each other within AAO pores with a mean diameter of 80

19682 J. Phys. Chem. C, Vol. 112, No. 49, 2008

Birjukovs et al.

Figure 4. Topography (left) and current maps (right) of vertically aligned 200 nm diameter Bi2S3 nanowires within AAO membranes.

nm. The corresponding fast Fourier transforms (FFT) of the HRTEM image (Figure 3a) verifies their single crystalline nature indicated by the sharp diffraction dots. The SAED patterns were taken with the electron beam along the [010] direction and confirmed that the nanowires were grown along the [002] direction. All of the nanowires studied by SAED and FFTs of the HRTEM images, isolated from the AAO, showed growth directions along the [002] zone axis. Detailed crystallography studies showed that this type of material is among the few examples of highly oriented and aligned single crystalline nanowires deposited as high density arrays within an electrically isolating matrix.29 Figure 4 shows the topography and current maps of the surface of vertically oriented Bi2S3 nanowires, with a mean diameter of 200 nm, measured through a 20 µm thick membrane. Ninety percent of the nanowires contacted using the vertical C-AFM approach were observed to be conductive, and these nanowires ran the entire length of the membrane. The resistivities of the Bi2S3 nanowires, with mean diameters of 200 and 80 nm, determined using this technique were found to be in the intervals 3-20 and 30-120 Ω m respectively. These resistivity measurements are orders of magnitude higher than the reported resistivity of bulk Bi2S3 (5 × 10-5 Ω · m),26 but they are in the range, or higher, compared to resistivity values obtained for individual Bi2S3 nanowires grown by templatefree methods and measured in a two-terminal configuration, for example, 0.06-12.130 and 0.012 Ω · m.31 We have previously shown that the ends of the nanowires can be damaged during the mechanical polishing process, thus being a source of enormous contact problems.18 In fact, the defects created at the ends of the nanowires can lead to nonuniform nanowire-probe interaction, increasing the resistances of nanowires to a large extent. In order to avoid contact to the damaged ends of nanowires, resistance measurments were performed on the longitudinal axes of the exposed nanowires. In this configuration, the conductive probe can contact a nanowire at different distances from the macroelectrode (Figure 2b). Moreover, it will allow a single

Figure 5. SEM images of a cross section of a Bi2S3/AAO matrix etched with 9% H3PO4 after (a) 0 min, (b) 120 min, (c) 180 min, and (d) 220 min.

nanowire to be scanned to perform a “quasi” multiprobe electrical measurement, thus allowing the contact resistance to be determined. Critical to these experiments is the degree of removal of the AAO matrix during the planarization process. As the grain size of the diamond pastes used for polishing were comparable to or larger than the diameters of the nanowires, most of the open nanowires were damaged during this process. Therefore, the samples were polished under angle and only small surface areas of nanowires were exposed (Figure 5a). Only areas containing intact nanowires were subjected to subsequent chemical etching of the AAO matrix with 9% H3PO4. The resistance measurements were conducted on the areas exposed by chemical etching. The length of the open ended parts of the nanowires could be further adjusted by varying the etching time in 9% H3PO4 (Figure 5b-d). For etching times longer than 120 min, completely open nanowires were obtained (Figure 5c,d). Since such nanowires are no longer isolated by the templating matrix, they can contact each other, thus the measured current when

Characterization of Bi2S3 Nanowire Arrays

J. Phys. Chem. C, Vol. 112, No. 49, 2008 19683

Figure 6. Cross-sectional conductive AFM images of Bi2S3 nanowires within AAO templates showing current maps ((a) 200 nm and (b) 80 nm). For (a), the distance from the macroelectrode was 12 µm, and for (b) -8 µm.

Figure 7. Current map of 200 nm diameter Bi2S3 nanowires within an AAO membrane and corresponding current profiles (b,c) across the nanowires as shown in (a).

Figure 8. I(V) characteristics of AAO-templated Bi2S3 nanowires measured with a doped diamond conductive AFM tip: (a) I(V) on 200 and 80 nm diameter nanowires; (b) I(V) on 200 nm diameter nanowires at different distances from the macroelectrode.

the AFM probe is not in contact to a single nanowire effectively reflects the current through a group of an unknown number of nanowires. Therefore, the resistivity was determined on samples with partly opened nanowires (see Figure 5b as an example) where the possibility of forming electrical contacts between neighboring nanowires is diminished. Figure 6 shows current maps of the surfaces of partially opened Bi2S3 nanowires with mean diameters of 200 and 80 nm. Dark areas correspond to the presence of open parts of nanowires on the surface. These nanowires are running parallel to each other, corresponding to orientation within the membrane as shown schematically in Figure 2b. Nanowires that were completely “open” were observed to be continuous between both contacts without breaks. The open lengths for 200 nm diameter nanowires were between 1-3 µm (Figure 6a), but those for 80 nm diameter wires were much shorter (Figure 6b). As shown in Figure 7, the current values are almost equally distributed across the nanowires. After exposure of the nanowires, the AFM probe was positioned on single nanowire and its I(V) charac-

teristics were obtained. The I(V) characteristics of individual Bi2S3 nanowires were nonlinear and exhibited a nonconductive gap for both the 200 and 80 nm diameter nanowires (Figure 8a). The nonconductive gap may be related to the presence of a thin amorphous layer on the nanowire surface31,32 and/or a Shottky barrier.33 The slope of the linear part of the I(V) curve at higher bias is related directly to the resistance of the nanowire.32,33 Figure 8b shows I(V) characteristics of a single 200 nm diameter nanowire measured along its length. The slope in the linear part of the I(V) curve is decreasing when the distance from the electrode is increasing and provides a direct indication that the contact resistance of a single nanowire can be readily extrapolated by performing length dependent measurements along the longitudinal axes of the nanowires. The contact resistance was determined by measuring the nanowire resistance at different probe-macroelectrode separations. The length of a nanowire was determined by scanning from the open end along the length of the nanowire until the contact point was reached, at which a sharp increase in the

19684 J. Phys. Chem. C, Vol. 112, No. 49, 2008

Birjukovs et al. Additionally, it should be noted that the resistance measurements obtained along the nanowire longitudinal axes are 10-100 times lower than the values obtained when the measurements were conducted with the vertical C-AFM configuration. These data therefore suggest that the horizontal C-AFM method used for contacting arrays of nanowires within AAO membranes can be used to estimate the “true” intrinsic resistivity of the nanowires and also present a valuable way of contacting aligned arrays of nanowires. Conclusions

Figure 9. Resistance of two, individual Bi2S3 nanowires within the AAO matrix measured by a single scan along the nanowire longitudinal axis with a doped diamond tip, as a function of the distance from the macroelectrode.

current was observed due to the metallic conductivity of the macroelectrode. Figure 9 shows that the resistance decreases as the separation between the probe and the macroelectrode is decreased, and becomes constant at separations below 1 µm. The increase in the nanowire resistance close to the bottom electrode may be due to the formation of a depletion layer in the nanowires. However, for Bi2S3 nanowires, the depletion distance is shorter, approximately 200 nm, than our electrode separation even at voltages 4 orders of magnitude higher than those used in the above measurements (2 mV).32 These data therefore suggest that the overall resistance (sum of nanowire and contact resistances) for probe-macroelectrode separations that are below 1 µm is determined by the contact resistance. The contact resistance may be determined by the presence of an amorphous oxide layer on the nanowire, the presence of an insulator layer on the AFM tip, or a Shottky barrier. Contact resistance depends on tip conditions and may change from sample to sample. For example, boron-doped diamond tips used on two different Bi2S3 nanowire samples yielded contact resistances of 15 and 3 MΩ (Figure 10). In some measurements with Pt-coated tips, the contact resistance reached values higher than 100 MΩ. We speculate that in these cases the Pt layer that covers the AFM tip is damaged and that most likely tunneling instead of an ohmic contact is occurring. For metal contacts deposited on individual semiconductor nanowires, the contact resistance is suggested to be around 1 MΩ.33 A factor that may play a substantial role in explaining the increased contact resistances in the C-AFM measurements, in comparison to measurements performed using deposited electrodes, is the difference in the size of the contacts. Effectively the contact area between the C-AFM probe and the nanowires is much smaller than the size of the deposited contact. This argument is supported by the fact that the contact resistivity and I(V) characteristics both depend on the size of the contact area as shown in recent papers.32,34 The resistivities of the Bi2S3 nanowires were determined by subtracting the contact resistance from the overall resistance. The resistivity for the 200 nm diameter nanowires was measured for approximately 470 nanowires on four different samples using two different AFM tips and was in the range between 0.4 and 1.2 Ω · m. The weighted average resistivity for the 80 nm diameter nanowires, measured for 200 nanowires on two different samples by using two different AFM probes, was 50 Ω · m. This is 1 order of magnitude higher than the resistivity measured for 200 nm diameter nanowires. The higher resistivity of 80 nm diameter nanowires can be explained by the greater influence of surface traps as seen with Ge nanowires.18

While the electrical properties of individual nanowires have been intensely investigated, there are relatively few studies of the conductive properties of ordered arrays of semiconductor nanowires within defined architectures. Until now, most of the investigations for determining the resistivity of nanowires within channeled matrices have been realized by contacting the exposed ends of vertically oriented nanowires. Alternatively, the other well investigated approach utilizes nanowires that are released from a matrix and contacted by fabricating metal contacts with focused ion beam or electron beam lithography. Here, we describe a novel method for contacting horizontal arrays of nanowires that has allowed us to access the full length of the nanowires. This technique employs precise polishing and selective etching steps to obtain exposed nanowires parallel to their longitudinal axes partially supported by the AAO membrane. The resistivities determined by this technique are 1-2 orders of magnitude lower than those measured by contacting the nanowires in a vertical configuration. Most importantly, we demonstrated that this technique can be applied to reliably determine the contact resistance by performing quasi multipoint measurements. Additionally, applying this contacting strategy, large numbers (more than several hundred) of aligned and oriented single crystal nanowires were electrically contacted and screened. In summary, this paper presents a viable method for electrically characterizing arrays of semiconductor nanowires deposited within templating matrices such as AAO. Acknowledgment. The authors acknowledge financial support from ERAF, Ministry of Education and Science of Latvia, the Centre for Research on Adaptive Nanostructures and Nanodevices (CRANN), and Science Foundation Ireland (Grant No. 03/IN3/I375). We acknowledge K. Didriksone for the help with sample preparation. References and Notes (1) Lieber, C. M. Sci. Am. 2001, 285, 50–56. (2) Cui, Y.; Lieber, C. M. Science 2001, 291, 851–853. (3) Wang, D.; Chang, Y.-L.; Wang, Q.; Cao, J.; Farmer, D. B.; Gordon, R. G.; Dai, H. J. Am. Chem. Soc. 2004, 126, 11602–11611. (4) Wang, D.; Wang, Q.; Javey, A.; Tu, R.; Dai, H.; Kim, H.; McIntyre, P. C.; Krishnamohan, T.; Saraswat, K. C. Appl. Phys. Lett. 2003, 83, 2432– 2434. (5) Duan, X. F.; Huang, Y.; Cui, Y.; Wang, J. E.; Lieber, C. M. Nature 2001, 409, 66–69. (6) Thelander, C.; Nilsson, H. A.; Jensen, L. E.; Samuelson, L. Nano Lett. 2005, 5, 635–638. (7) Ziegler, K. J.; Lyons, D. M.; Holmes, J. D.; Erts, D.; Polyakov, B.; Olin, H.; Svensson, K.; Olsson, E. Appl. Phys. Lett. 2004, 84, 4074– 4076. (8) Cui, Y.; Wei, Q. Q.; Park, H.; Lieber, C. M. Science 2001, 293, 1289–1292. (9) Huang, Y.; Duan, X.; Cui, Y.; Lauhon, L. J.; Kim, K. H.; Lieber, C. M. Science 2001, 294, 1313–1317. (10) Ziegler, K.; Ryan, K. M.; Rice, R.; Crowley, T. A.; Erts, D.; Olin, H.; Patterson, J.; Spalding, T. R.; Holmes, J. D.; Morris, M. A. Faraday Discuss. 2004, 125, 311–326. (11) Lew, K. K.; Redwing, J. M. J. Cryst. Growth 2003, 254, 14–22.

Characterization of Bi2S3 Nanowire Arrays (12) Liang, Y.; Zhen, C.; Zou, D.; Xu, D. J. Am. Chem. Soc. 2004, 126, 16338–16339. (13) Xu, J.; Petkov, N.; Wu, X.; Iacopino, D.; Quinn, A. J.; Redmond, G.; Bein, T.; Morris, M. A.; Holmes, J. D. ChemPhysChem 2007, 8, 235– 240. (14) Peng, X. S.; Meng, G. W.; Zhang, J.; Zhao, L. X.; Wang, X. F.; Wang, Y. W.; Zhang, L. D. J. Phys. D: Appl. Phys. 2001, 34, 3224–3228. (15) Ryan, K. M.; Erts, D.; Olin, H.; Morris, M. A.; Holmes, J. D. J. Am. Chem. Soc. 2003, 125, 6284–6889. (16) Liu, C. H.; Yiu, W. C.; Au, F. C. K.; Ding, J. X.; Lee, C. S.; Lee, S. T. Appl. Phys. Lett. 2003, 83, 3168–3170. (17) Ziegler, K. J.; Polyakov, B.; Kulkarni, J. S.; Crowley, T. A.; Ryan, K. M.; Morris, M. A.; Erts, D.; Holmes, J. D. J. Mater. Chem. 2004, 14, 585–589. (18) Erts, D.; Polyakov, B.; Daly, B.; Morris, M. A.; Ellingboe, S.; Boland, J.; Holmes, J. D. J. Phys. Chem. B 2006, 110, 820–826. (19) Polyakov, B.; Daly, B.; Prikulis, J.; Lisauskas, V.; Vengalis, B.; Holmes, J. D.; Erts, D. AdV. Mater. 2006, 18, 1812–1816. (20) Petkov, N.; Birjukovs, P.; Phelan, R.; Morris, M. A.; Erts, D.; Holmes, J. D. Chem. Mater. 2008, 20, 1902–1908. (21) Dayen, J.-F.; Rumyantseva, A.; Ciornei, C.; Wade, T. L.; Wegrowe, J.-E.; Pribat, D.; Sorin Cojocaru, C. Appl. Phys. Lett. 2007, 90, 173110– 173113. (22) Mohammad, A. M.; Dey, S.; Lew, K.-K.; Redwing, J. M.; Mohney, S. E. J. Electrochem. Soc. 2003, 150, 577–580.

J. Phys. Chem. C, Vol. 112, No. 49, 2008 19685 (23) Eichfeld, S. M.; Ho, T.-T.; Eichfeld, C. M.; Cranmer, A.; Mohney, S. E.; Mayer, T. S.; Redwing, J. M. Nanotechnology 2007, 18, 315201– 315206. (24) Pawar, S. H.; Tamhankar, S. P.; Lokhande, C. D. Mater. Chem. Phys. 1984, 11, 401–412. (25) Mane, R. S.; Sankapal, B. R.; Lokhande, C. D. Mater. Chem. Phys. 1999, 60, 158–162. (26) Chen, B. X.; Uher, C.; Iordanidis, L.; Kanatzidis, M. G. Chem. Mater. 1997, 9, 1655–1658. (27) Masuda, H.; Fukuda, K. Science 1995, 268, 1466–1468. (28) Masuda, H.; Yamada, H.; Satoh, M.; Asoh, H.; Nakao, M. Appl. Phys. Lett. 1997, 71, 2770–2772. (29) Petkov, N.; Xu, J.; Morris, M. A.; Holmes, J. D. J. Phys. Chem. C 2008, 112, 7345–7355. (30) Schricker, A.; Sigman, M., Jr.; Korgel, B. Nanotechnology 2005, 16, 508–513. (31) Yu, Y.; Jin, C. H.; Wang, R. H.; Chen, Q.; Peng, L.-M. J. Phys. Chem. B 2005, 109, 18772–18776. (32) Zhang, Z.; Yao, K.; Liu, Y.; Liang, X.; Chen, Q.; Peng, L.-M. AdV. Funct. Mater. 2007, 17, 24782489. (33) Zheng, G.; Lu, W.; Jin, S.; Lieber, C. M. AdV. Mater. 2006, 16, 1890–1893. (34) Chaudhry, A.; Ramamurthi, V.; Fong, E.; Islam, M. S. Nano Lett. 2007, 7, 1536–1541.

JP805422K